首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 0 毫秒
1.
Liquid–(solid + liquid) transitions are studied in (CH3)CCl3 + CCl4 by using the Landau phenomelogical model. The Gibbs energy is expanded in terms of the orientational disorder (OD) parameters for the transitions of the liquid–(rhombohedral + liquid) and liquid–(face-centered cubic + liquid) in a two component system of (CH3)CCl3 + CCl4. From the Gibbs energy, the phase line equations are derived for the transitions studied and they are fitted to the observed T–X phase diagram of (CH3)CCl3 + CCl4 for the concentration (X) CCl4. Temperature and concentration dependences of the OD parameters (Ψ and η) and the inverse susceptibility ($$\chi_{\psi }^{ - 1}$$ and $$\chi_{\eta }^{ - 1}$$) for the two transitions of interest, are predicted by using the melting curves of (CH3)CCl3 + CCl4 on the basis of the Landau phenomenological model. Our predictions, which can be compared with the experimental data, indicate that the first order transition of the liquid–(solid + liquid), in particular, for (CH3)CCl3 + CCl4 can be described satisfactorily by the Landau mean field model.  相似文献   

2.
Liquid–liquid equilibria (LLE) data for {water + 1-butanol + isophorone} and {water + 1-butanol + mesityl oxide} ternary systems were investigated systematically at different temperatures under atmospheric pressure. The Othmer–Tobias and Bachman equations were applied to analyze the dependability of the experimental LLE data. Selectivities and distribution coefficients were used to evaluate the extractive efficiency of the extractants. The Non-Random Two Liquid (NRTL) and Universal Quasi-Chemical (UNIQUAC) models were applied to correlate the studied systems and were well represented with all root mean square deviations (RMSD) less than 0.2%. Meanwhile, binary interaction parameters among these compounds were acquired during the correlation process.  相似文献   

3.
4.
Densities, $$\rho$$, and kinematic viscosities, $$\nu$$, have been determined at atmospheric pressure and at 293.15–303.15 K for binary mixtures formed by methanol and one linear polyether of the type CH3–O–(CH2CH2O)n–CH3 (n = 2, 3, 4). Measurements on $$\rho$$ and $$\nu$$ were carried out, respectively, using an Anton Paar DMA 602 vibrating-tube densimeter and an Ubbelohde viscosimeter. The $$\rho$$ values were used to compute excess molar volumes, $$V_{{\text{m}}}^{{\text{E}}}$$, and, together with the $$\nu$$ results, dynamic viscosities ($$\eta$$). Deviations from linear dependence on mole fraction for viscosity, $$\Delta \eta$$, are also provided. Different semi-empirical equations have been employed to correlate viscosity data. Particularly, the equations used are the: Grunberg–Nissan, Hind, Frenkel, Katti–Chaudhri, McAllister and Heric. Calculations show that better results are obtained from the Hind equation. The $$V_{{\text{m}}}^{{\text{E}}}$$ values are large and negative and contrast with the positive excess molar enthalpies, $$H_{{\text{m}}}^{{\text{E}}}$$, available in the literature, for these systems. This indicates that structural effects are dominant. The $$\Delta \eta$$ results are positive and correlate well with the difference in volumes of the mixture compounds, confirming the importance of structural effects. The temperature dependences of $$\eta$$ and of the molar volume have been used to calculate enthalpies, entropies and Gibbs energies, $$\Delta G^{*}$$, of viscous flow. It is demonstrated that $$\Delta G^{*}$$ is essentially determined by enthalpic effects. Methanol + CH3–O–(CH2CH2O)n–CH3 mixtures have been treated in the framework of the ERAS model. Results for $$H_{{\text{m}}}^{{\text{E}}}$$ are acceptable, while the composition dependence of the $$V_{{\text{m}}}^{{\text{E}}}$$ curves is poorly represented. This has been ascribed to the existence of strong dipolar and structural effects in the present solutions.  相似文献   

5.
The activation Gibbs energies, activation enthalpies and activation entropies reported by the authors were found to be internally inconsistent. Errors were also found in the numerical values of the Redlich–Kister equation coefficients for describing the excess molar volumes and viscosity deviations of both binary systems studied by the authors.  相似文献   

6.
Solid-liquid phase equilibria of the two ternary systems (NaCl+SrCl2+H2O) and (KCl+SrCl2+H2O) at T=288.15 K and p=0.1 MPa were studied using the isothermal dissolution equilibrium method. Solubilities of the equilibrium liquid phase were determined, and the solids were also investigated by the Schreinemaker method of wet residues. In the ternary system (NaCl+SrCl2+H2O) at 288.15 K, there is one invariant point corresponding to (NaCl+SrCl2·6H2O) and two crystallization regions corresponding to NaCl and SrCl2·6H2O. The crystallized area of SrCl2·6H2O decreased with the increasing temperature, while that of NaCl increased slightly. In the ternary system (KCl+SrCl2+H2O) at 288.15 K, there is one invariant point(KCl+SrCl2·6H2O) and two crystallization regions corresponding to KCl and SrCl2·6H2O. Both systems belong to a simple eutectic type, and neither double salts nor solid solutions were formed. On the basis of Pitzer-Harvie-Weare model, the solubilities of the two systems at 288.15 K were demonstrated. A comparison showed that the calculated solubilities agreed well with the experimental data.  相似文献   

7.
8.
9.
10.
Metastable equilibrium solubilities and properties such as densities, conductivity, pH, refractive index, and viscosity of the solution were determined experimentally. According to the experimental data, the metastable equilibrium phase diagram was plotted. In the phase diagram, there are three invariant points, seven univariant curves, five fields of crystallization: Li2SO4 · H2O, K2SO4, Li2B4O7 · 3H2O, K2B4O7 · 4H2O, and K2SO4 · Li2SO4. The double salt K2SO4 · Li2SO4 was found in the quaternary system metastable equilibria. Lithium sulfate (Li2SO4) has the highest concentration and strong salting-out effects on other salts.Also, the relationship diagram between the properties and the ion concentration of solution was constructed. It can be seen from the relationship diagram that the equilibrium solution density values, viscosity values, and refractive index values are increased apparently with the rise of sulfate ion concentration, reaching the maximum values at eutonic point F3. Electrical conductivity values and pH values, however, fall down with the rise of ion concentration on the whole.  相似文献   

11.
钽酸盐光催化材料往往具有较高的光催化活性.近年报道的钽酸盐光催化剂主要采用传统高温固相法制备,该方法不可避免地导致高温烧结,使合成的钽酸盐颗粒较大,比表面积较小,而且该方法具有不可克服的晶体转变、结晶度差、分解、挥发和纯度低等缺点,使制备的光催化剂活性较低.而纳米材料由于粒径小,提高了电子和空穴的扩散速度,大大降低了电子和空穴在材料内的复合几率,从而使光催化材料活性大幅提高.此外,粒径减小也使表面原子迅速增多,减小了光的漫反射,同时也使光吸收不易达到饱和,有利于提高光吸收效率.因此,制备纳米材料是提高半导体光催化剂活性的有效手段.目前,采用湿化学的溶液合成方法能在较低温度下获得粒度小且均匀、计量比准确的光催化剂粉末,但是合成钽酸盐光催化剂的水溶性钽前体即乙醇钽(或氯化钽)价格昂贵,而且对潮湿极端敏感易水解,使产物纯度降低,不适合工业化生产.近年来,尽管有文献报道以Ta2O5为原料利用水热、溶胶-凝胶和共沉淀等方法制备钽酸盐,但其合成条件苛刻,合成步骤复杂,合成周期较长,耗能大,产物产量较低且不均匀,很难实现产物的形貌控制来筛选出适合光催化反应的材料.目前关于纳米钽酸盐光催化材料形貌控制方面的研究鲜有报道,主要是由于Ta2O5极难溶解,很难实现液相合成.因此,纳米钽酸盐光催化材料的可控制备是研究的难点.我们发展了熔盐-水热制备钽酸盐新方法,实现了K1.9Na0.1Ta2O6·2H2O的可控制备.利用熔盐法制备一种可溶性钽酸盐前驱体,再通过水热法在液相进一步反应制得纳米钽酸盐光催化材料K1.9Na0.1Ta2O6·2H2O,通过控制反应条件实现了纳米钽酸盐K1.9Na0.1Ta2O6·2H2O的形貌调控,得到了纳米球、微球、去顶八面体形貌和类似榴莲形貌等不同形貌,而利用其它制备方法很难控制钽酸盐的形貌.另外,研究了制备材料吸附和光催化降解罗丹明B的性能,发现该材料光催化活性与形貌直接相关.表征结果表明,制备样品的X射线衍射(XRD)谱图尖锐,结晶较好,其各衍射峰位置均与K2Ta2O6一致,为纯相烧绿石结构,属于立方晶系,空间群为Fd3m.通过分析合成材料的元素组成及含量,确定K:Na:Ta比例近似为1.9:0.1:2.为了进一步研究属于烧绿石型化合物K1.9Na0.1Ta2O6·2H2O的结构,对不同形貌材料进行了红外光谱测试,所有样品在450–1000 cm–1的谱峰可归属于(K, Na)–O和Ta–O键的振动,3300 cm–1左右为晶体结构中水的羟基伸缩振动峰,1720 cm–1左右是晶体结构中水的弯曲振动峰.可以看出,不同形貌材料的红外谱图吸收带宽度和位置十分相似,只存在小的偏移和变化,进一步表明不同形貌的材料具有相似的晶体结构,与XRD结果一致.差热-热重分析确定了结构中所含结晶水数量近似为2.光催化性能测试结果表明,具有纳米球形貌的材料比表面积较大,因而光催化活性最高.  相似文献   

12.
The equilibrium solubility of the quaternary system CsBr-CeBr3-HBr(12.52 wt %)-H2O was determined at 298K and the corresponding equilibrium diagram was constructed. The quaternary system is complicated by three equilibrium solid phases: CsBr, Cs5Ce2Br11 · 22H2O (5: 2 type), and CeBr3 · 7H2O, of which the new compound Cs5Ce2Br11 · 22H2O was found to be incongruently soluble in the system. The new compound obtained was identified and characterized by the method of X-ray diffraction and the thermal analysis methods of thermogravimetry-differential thermogravimetry (TG-DTG), and it loses its crystal water in two steps from 325 to 511 K. The standard molar enthalpy of solution of Cs5Ce2Br11 · 22H2O in deionized water was measured to be (129.105 ± 0.150) kJ mol−1 with a heat conduction microcalorimeter. The standard molar enthalpy of formation was calculated as (−10438.215 ± 0.150) kJ mol−1. This article was submitted by the authors in English.  相似文献   

13.
14.
15.
Binary and ternary parameters of the Pitzer equation and also the thermodynamic potentials of solid phases and the solubility diagrams of ternary aqueous-salt systems Na+, K+||Cl--H2O and Na+, K+||Cl-, H2PO- 4-H2O in the temperature range 0-100°C, which corresponds to the technological conditions of brine formation, were calculated.  相似文献   

16.
Experimental tie-line data were determined for one ternary system, water + diisopropyl ether + n-heptane and two quaternary systems, water + diisopropyl ether + 2-propanol + n-heptane or toluene at 298.15 K and ambient pressure. The experimental liquid–liquid equilibrium data were successfully correlated using a modified UNIQUAC model with ternary and quaternary mixture parameters, in addition to the binary ones. The calculated results were also compared with those obtained from an extended UNIQUAC model of Nagata [Fluid Phase Equilib. 54, 191 (1990)].  相似文献   

17.
The metastable solubilities and the physicochemical properties including density and pH of the reciprocal quaternary system(LiCl+MgCl2+Li2SO4+MgSO4+H2O) at 348.15 K and 0.1 MPa were determined using the isothermal evaporation method. The dry-salt diagram and water-phase diagram were plotted based on the experimental data. There are five invariant points, eleven univariant curves, and seven crystallization zones corresponding to hexahydrite, tetrahydrite, kieserite, bischofite, lithium sulfate monohydrate, lithium chloride monohydrate and lithium carnallite. Comparison between the stable and metastable diagrams at 348.15 K indicates that the metastable phenomenon of magnesium sulfate is obvious, and the crystallization regions of hexahydrite and tetrahydrite disappear in the stable phase diagram. A comparison of the metastable dry-salt phase diagrams at 308.15, 323.15 and 348.15 K shows that with the increasing of temperature the epsomite crystallization zone disappears from the dry-salt phase diagram of 303.15 K, and a new kieserite crystallization zone is presented at 348.15 K. The density and pH in the metastable equilibrium solution present regular change with the increasing of Jänecke index J(2Li+), and the calculated densities using the empirical equation agree well with the experimental values.  相似文献   

18.
An experimental study on metastable equilibria at T=288 K in the quinary system Li2CO3 + Na2CO3 + K2CO3 + Li2B4O7 + Na2B4O7 + K2B4O7 + H2O was done by isothermal evaporation method. Metastable equilibrium solubilities and densities of the solution were determined experimentally. According to the experimental data, the metastable equilibrium phase diagram under the condition saturated with Li2CO3 was plotted, in which there are four invariant points; nine univariant curves; six fields of crystallization: K2CO3 · 3/2H2O, K2B4O7 · 5H2O, Li2B2O4 · 16H2O, Na2B2O4 · 8H2O, Na2CO3 · 10H2O, NaKCO3 · 6H2O. Some differences were found between the stable phase diagram at T=298 K and the metastable one at T=288 K.  相似文献   

19.
New isothermal pTxy data are reported for (methane + benzene) and (methane + methylbenzene (toluene)) at pressures up to 13 MPa over the temperature range (188 to 313) K using a custom-built (vapor + liquid) equilibrium (VLE) apparatus. The aim of this work was to investigate literature data inconsistencies and to extend the measurements to lower temperatures. For (methane (1) + benzene (2)), measurements were made along six isotherms from (233 to 348) K at pressures to 9.6 MPa. At temperatures below 279 K there was evidence of a solid phase, and thus only vapor phase samples were analyzed at these temperatures. For the (methane (1) + methylbenzene (3)) system, measurements were made along seven isotherms from T = (188 to 313) K at pressures up to 13 MPa. Along the 198 K isotherm, a significant change in the data’s p,x slope was observed indicating (liquid + liquid) equilibria at higher pressures. The data were compared with literature data and with calculations made using the Peng–Robinson (PR) equation of state (EOS). For both binary systems our data agree with much of the literature data that also deviate from the EOS in a similar manner. However, the data of Elbishlawi and Spencer (1951) for both binary systems, which appear to have received an equal weighting to other data in the EOS development, are inconsistent with the results of our measurements and data from other literature sources.  相似文献   

20.
The excess molar volumes and the partial molar volumes for (propionitrile + an alkanol) at T = 298.15 K and at atmospheric pressure are reported. The hydrogen bonding between the OH⋯NC groups are discussed in terms of the chain length of the alkanol. The alkanols studied are (methanol, ethanol, 1-propanol, 2-propanol, 1-butanol, and 1-pentanol).The excess molar volume data was fitted to the Redlich–Kister equation The partial molar volumes were calculated from the Redlich–Kister coefficients.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号